You are on page 1of 25

Chapter 14

Thin airfoil theory

14.1 Compressible potential flow

14.1.1 The full potential equation

In compressible flow, both the lift and drag of a thin airfoil can be determined to a reason-
able level of accuracy from an inviscid, irrotational model of the flow. Recall the equations
developed in Chapter 6 governing steady, irrotational, homentropic (rs = 0) flow in the
absence of body forces.

r U = 0

U U rP
r + =0
2 (14.1)

P
=
P0 0

The gradient of the isentropic relation is

rP = a2 r. (14.2)

Recall from the development in Chapter 6 that



P 1 rP
r = . (14.3)

14-1
CHAPTER 14. THIN AIRFOIL THEORY 14-2

Using (14.3) the momentum equation becomes


P U U
r + = 0. (14.4)
1 2

Substitute (14.2) into the continuity equation and use (14.3). The continuity equation
becomes

U ra2 + ( 1) a2 r U = 0. (14.5)

Equate the Bernoulli integral to free stream conditions.


a2 U U a1 2 U1 2 a1 2 1 2
+ = + = 1+ M1 = C p Tt (14.6)
1 2 1 2 1 2

Note that the momentum equation is essentially equivalent to the statement that the
stagnation temperature Tt is constant throughout the flow. Using (14.6) we can write

a2 U U
= ht . (14.7)
1 2

The continuity equation finally becomes


U U U U
( 1) ht r U U r = 0. (14.8)
2 2

The equations governing compressible, steady, inviscid, irrotational motion reduce to a


single equation for the velocity vector U. The irrotationality condition r U = 0 permits
the introduction of a velocity potential.

U = r (14.9)

and (14.8) becomes



r r 2 r r
( 1) ht r r r = 0. (14.10)
2 2

For complex body shapes numerical methods are normally used to solve for . However
the equation is of relatively limited applicability. If the flow is over a thick airfoil or a blu
CHAPTER 14. THIN AIRFOIL THEORY 14-3

body for instance then the equation only applies to the subsonic Mach number regime at
Mach numbers below the range where shocks begin to appear on the body. At high subsonic
and supersonic Mach numbers where there are shocks then the homentropic assumption
(14.2) breaks down. Equation (14.8) also applies to internal flows without shocks such as
fully expanded nozzle flow.

14.1.2 The nonlinear small disturbance approximation

In the case of a thin airfoil that only slightly disturbs the flow, equation (14.8) can be
simplified using small disturbance theory. Consider the flow past a thin 3-D airfoil shown
in Figure 14.1.

Figure 14.1: Flow past a thin 3-D airfoil

The velocity field consists of a free-stream flow plus a small disturbance

U = U1 + u

V =v (14.11)

W =w

where

u
<< 1
U1
v
<< 1 (14.12)
U1
w
<< 1.
U1

Similarly the state variables deviate only slightly from free-stream values
CHAPTER 14. THIN AIRFOIL THEORY 14-4

P = P1 + P 0

T = T1 + T 0 (14.13)

= 1 + 0

and

a = a1 + a0 . (14.14)

This decomposition of variables is substituted into equation (??)(eq:original014008). Var-


ious terms are

U U U1 2 u2 v 2 w 2
= + uU1 + + +
2 2 2 2 2 (14.15)
r U = ux + vy + wz

and


U U
r = (ux U1 + uux + vvx + wwx , uy U1 + uuy + vvy + wwy , uz U1 + uuz + vvz + wwz )
2
(14.16)
as well as


U U
( 1) ht r U =
2

U1 2 u2 v 2 w 2
( 1) ht + uU1 + + + ux +
2 2 2 2
(14.17)
U1 2 u2 v 2 w 2
( 1) ht + uU1 + + + uy +
2 2 2 2

U1 2 u2 v 2 w 2
( 1) ht + uU1 + + + uz
2 2 2 2

and, finally
CHAPTER 14. THIN AIRFOIL THEORY 14-5


U U U U
( 1) ht r U U r =
2 2

U1 2
( 1) ht + uU1 ux +
2

U1 2
( 1) ht + uU1 vy +
2 (14.18)

U1 2
( 1) ht + uU1 wz
2

ux U1 2 + uux U1 + vvx U1 + wwx U1

uux U1 vuy U1 wuz U1 .

Neglect terms in (14.17) and (14.18) that are of third order in the disturbance velocities.
Now


U U U U
( 1) ht r U U r =
2 2

U1 2
( 1) ht + uU1 ux +
2

U1 2
( 1) ht + uU1 vy +
2 (14.19)

U1 2
( 1) ht + uU1 wz
2

ux U1 2 + uux U1 + vvx U1 + wwx U1

uux U1 vuy U1 wuz U1

or
CHAPTER 14. THIN AIRFOIL THEORY 14-6


U U U U
( 1) ht r U U r =
2 2

( 1) (h1 uU1 ) (ux + vy + wz )


(14.20)
ux U1 2 + uux U1 + vvx U1 + wwx U1

uux U1 vuy U1 wuz U1 .

Recall that ( 1) h1 = a1 2 . Equation (14.20) can be rearranged to read


U U U U
( 1) ht r U U r =
2 2
(14.21)
a1 2 (ux + vy + wz ) ux U 1 2 ( + 1) uux U1 (vvx U1 + wwx U1 )

( 1) (uvy U1 + uwz U1 ) vuy U1 wuz U1 .

Divide through by a1 2 .


1 U U 1 U U
ht r U U r =
a1 2 2 a1 2 2

( + 1) M1
1 M1 2 u x + v y + w z uux (14.22)
a1

M1
(( 1) (uvy + uwz ) + vuy + wuz + vvx + wwx )
a1

Equation (14.22) contains both linear and quadratic terms in the velocity disturbances
and one might expect to be able to neglect the quadratic terms. But note that the first
term becomes very small near M1 = 1. Thus in order to maintain the small disturbance
approximation at transonic Mach numbers the uux term must be retained. The remaining
quadratic terms are small at all Mach numbers and can be dropped. Finally the small
disturbance equation is

( + 1) M1
1 M1 2 u x + v y + w z uux = 0. (14.23)
a1
CHAPTER 14. THIN AIRFOIL THEORY 14-7

The velocity potential is written in terms of a free-stream potential and a disturbance


potential.

= U1 x + (x, y, z) (14.24)

The small disturbance equation in terms of the disturbance potential is

( + 1) M1
1 M1 2 xx + yy + zz = x xx . (14.25)
a1

Equation (14.25) is valid over the whole range of subsonic, transonic and supersonic Mach
numbers.

14.1.3 Linearized potential flow

If we restrict our attention to subsonic and supersonic flow, staying away from Mach
numbers close to one, the nonlinear term on the right side of (14.25) can be dropped and
the small disturbance potential equation reduces to the linear wave equation

2
xx ( yy + zz ) =0 (14.26)
q
where = M1 2 1. In two dimensions

2
xx yy = 0. (14.27)

The general solution of (14.27) can be expressed as a sum of two arbitrary functions

(x, y) = F (x y) + G (x + y) . (14.28)

Note that if M1 2 < 1 the 2-D linearized potential equation (14.27) is an elliptic equation
that can be rescaled to form Laplaces equation and (14.28) expresses the solution in terms
of conjugate complex variables. In this case the subsonic flow can be analyzed using the
methods of complex analysis. Presently we will restrict our attention to the supersonic
case. The subsonic case is treated later in the chapter.
If M1 2 > 1 then (14.27) is the 2-D wave equation and has solutions of hyperbolic type.
Supersonic flow is analyzed using the fact that the properties of the flow are constant along
the characteristic lines x y = constant. Figure 14.2 illustrates supersonic flow past a
CHAPTER 14. THIN AIRFOIL THEORY 14-8

thin airfoil with several characteristics shown. Notice that in the linear approximation the
characteristics are all parallel to one another and lie at the Mach angle 1 of the free
stream. Information about the flow is carried in the value of the potential assigned to
a given characteristic and in the spacing between characteristics for a given flow change.
Right-leaning characteristics carry the information about the flow on the upper surface
of the wing and left-leaning characteristics carry information about the flow on the lower
surface.

Figure 14.2: Right and left leaning characteristics on a thin 2-D airfoil.

All properties of the flow, velocity, pressure, temperature, etc. are constant along the char-
acteristics. Since disturbances only propagate along downstream running characteristics
we can write the velocity potential for the upper and lower surfaces as

(x, y) = F (x y) y > 0
(14.29)
(x, y) = G (x + y) y < 0.

Let y = f (x) define the coordinates of the upper surface of the wing and y = g (x) define
the lower surface. The full nonlinear boundary condition on the upper surface is

v df
= . (14.30)
U y=f dx

In the spirit of the thin airfoil approximation this boundary condition can be approximated
by the linearized form

v df
= (14.31)
U y=0 dx

which we can write as


CHAPTER 14. THIN AIRFOIL THEORY 14-9

@ (x, y) df
= U1 (14.32)
@y y=0 dx

or

U1 df
F 0 (x) = . (14.33)
dx

On the lower surface the boundary condition is

U1 dg
G0 (x) = . (14.34)
dx

In the thin airfoil approximation the airfoil itself is, in eect, collapsed to a line along
the x -axis, the velocity potential is extended to the line y = 0 and the surface boundary
condition is applied at y = 0 instead of at the physical airfoil surface. The entire eect of
the airfoil on the flow is accounted for by the vertical velocity perturbation generated by
the local slope of the wing. The linearized boundary condition is valid on 2-D thin wings
and on 3-D wings of that are of thin planar form.
Recall that (14.26) is only valid for subsonic and supersonic flow and not for transonic flow
where 1 M1 2 << 1.

14.1.4 The pressure coefficient

Lets work out the linearized pressure coefficient. The pressure coefficient is

P P1 2 P
CP = 1 2 = 2 1 . (14.35)
U
2 1 1 M 1 P1

The stagnation enthalpy is constant throughout the flow. Thus

T 1
=1+ U1 2 U 2 + v 2 + w2 . (14.36)
T1 2Cp T1

Similarly the entropy is constant and so the pressure and temperature are related by


P 1 1
= 1+ U1 2 2
U +v +w 2 2
. (14.37)
P1 2Cp T1
CHAPTER 14. THIN AIRFOIL THEORY 14-10

The pressure coefficient is

!
2 1 1
CP = 1+ U1 2 U 2 + v 2 + w2 1 . (14.38)
M1 2 2Cp T1

The velocity term in (14.37) is small.

U1 2 U 2 + v 2 + w2 = 2uU1 + u2 + v 2 + w2 (14.39)

Now use the binomial expansion (1 ")n = 1 n" + n (n 1) "2 /2 to expand the term in
parentheses in (14.39). Note that the expansion has to be carried out to second order. The
pressure coefficient is approximately


2u u2 v 2 + w2
CP
= + 1 M1 2
2 + . (14.40)
U1 U1 U1 2

Equation (14.40) is a valid approximation for small perturbations in subsonic or supersonic


flow.
For 2-D flows over planar bodies it is sufficient to retain only the first term in (14.40) and
we use the expression

u
CP
= 2 . (14.41)
U1

For 3-D flows over slender approximately axisymmetric, bodies we must retain the last
term so


2u v 2 + w2
CP
= + . . (14.42)
U1 U1 2

As was discussed above, if the airfoil is a 2-D shape defined by the function y = f (x) the
boundary condition at the surface is

df v
= = tan () (14.43)
dx U1 + u

where is the angle between the airfoil surface and the horizontal. For a thin airfoil this
is accurately approximated by
CHAPTER 14. THIN AIRFOIL THEORY 14-11

df y
= = . (14.44)
dx U1

For a thin airfoil in supersonic flow the wall pressure coefficient is



2 df
CPwall = 1/2
(14.45)
M1 2 1 dx
q
2 2
where dU /U = 2/ M1 2 1 d, which is valid for M1 1 has been used. In the
thin airfoil approximation in supersonic flow the local pressure coefficient is determined by
the local slope of the wing. Figure 14.3 schematically shows the wall pressure coefficient
on a thin, symmetric biconvex wing.

Figure 14.3: Pressure distribution on a thin symmetric airfoil.

This airfoil will have shock waves at the leading and trailing edges and at first sight this
would seem to violate the isentropic assumption. But for small disturbances the shocks
are weak and the entropy changes are negligible.

14.1.5 Drag coefficient of a thin symmetric airfoil

A thin, 2-D, symmetric airfoil is situated in a supersonic stream at Mach number M1 and
zero angle of attack. The y-coordinate of the upper surface of the airfoil is given by the
function
x
y (x) = ASin (14.46)
C
where C is the airfoil chord. The airfoil thickness to chord ratio is small, 2A/C << 1.
Determine the drag coefficient of the airfoil.
Solution
CHAPTER 14. THIN AIRFOIL THEORY 14-12

The drag integral is

Z C
D=2 (P P1 )Sin () dx. (14.47)
0

where the factor of 2 accounts for the drag of both the upper and lower surfaces and is
the local angle formed by the upper surface tangent to the airfoil and the x-axis. Since the
airfoil is thin the angle is small and we can write the drag coefficient as

Z !
D 1
P P1 x
CD = 1 2 =2 1 2 () d . (14.48)
2 1 U1 C
0 2 1 U1
C

The local tangent is determined by the local slope of the airfoil therefore T an () = dy/dx
and for small angles = dy/dx. Now the drag coefficient is

Z !
1
D P P1 dy x
CD = 1 2 =2 1 2 d . (14.49)
2 1 U1 C 0 2 1 U1
dx C

The pressure coefficient on the airfoil is given by thin airfoil theory (14.45) as

P P1 2 dy
CP = 1 2 =
p (14.50)
2 1 U1 M1 2 1 dx

and so the drag coefficient becomes

Z 1 2 Z 1 x x
4 dy x 4A2 2A2 2
CD = p d = p Cos2 d = p .
M1 2 1 0 dx C C 2 M1 2 1 0 C C C 2 M1 2 1
(14.51)
The drag coefficient is proportional to the square of the wing thickness-to-chord ratio.

14.1.6 Thin airfoil with lift and camber at a small angle of attack

A thin, cambered, 2-D, airfoil is situated in a supersonic stream at Mach number M1 and
a small angle of attack as shown in Figure 14.4.
The y-coordinate of the upper surface of the airfoil is given by the function
CHAPTER 14. THIN AIRFOIL THEORY 14-13

Figure 14.4: Trajectory of a piston and an expansion wave in the x-t plane.

x x x x
f = A +B (14.52)
C C C C

and the y-coordinate of the lower surface is

x x x x
g = A +B (14.53)
C C C C

where 2A/C << 1 and /C << 1. The airfoil surface is defined by a dimensionless
thickness function
x

C
(14.54)
(0) = (1) = 0

a dimensionless camber function


x
C
(14.55)
(0) = (1) = 0

and the angle of attack

T an () = . (14.56)
C

Determine the lift and drag coefficients of the airfoil. Let

x
= . (14.57)
C

Solution
CHAPTER 14. THIN AIRFOIL THEORY 14-14

The lift integral is

Z C Z C
L= (Plower P1 )Cos (lower ) dx (Pupper P1 )Cos (upper ) dx. (14.58)
0 0

where is the local angle formed by the tangent to the surface of the airfoil and the x-axis.
Since the angle is everywhere small we can use Cos () = 1. The lift coefficient is

Z 1 Z 1
D
CL = 1 2 = CPlower d CPupper d. (14.59)
2 1 U1 C 0 0

The pressure coefficient on the upper surface is given by thin airfoil theory as

2 df 2 d d
CPupper = p = p A +B . (14.60)
C M1 2 1 d C M1 2 1 d d

On the lower surface the pressure coefficient is



2 dg 2 d d
CPlower = p = p A +B . (14.61)
C M1 2 1 d C M1 2 1 d d

Substitute into the lift integral

Z 1 Z 1 Z Z
2 df dg 2
CL = p d + d = p df + dg .
C M1 2 1 0 d 0 d C M1 2 1 0 0
(14.62)
In the thin airfoil approximation, the lift is independent of the airfoil thickness.

4
CL = p (14.63)
M1 2 1 C

The drag integral is

Z C Z C
D= (Pupper P1 )Sin (upper ) dx + (Plower P1 )Sin ( lower ) dx. (14.64)
0 0
CHAPTER 14. THIN AIRFOIL THEORY 14-15

Since the airfoil is thin the angle is small, and we can write the drag coefficient as

Z ! Z !
1 1
D Pupper P1 Plower P1
CD = 1 2 = 1 2 (upper ) d + 1 2 ( lower ) d.
2 1 U 1 C 0 2 1 U1 0 2 1 U1
(14.65)
The local tangent is determined by the local slope of the airfoil therefore T an () = dy/dx
and for small angles = dy/dx. Now the drag coefficient is

Z 1 Z 1
D 1 dyupper 1 dylower
CD = 1 2 = CPupper d + CPlower d. (14.66)
2 1 U 1 C
C 0 d C 0 d

The drag coefficient becomes

!
Z 1 2 Z 1 2
2 d d d d
CD
= p A +B d + A +B d .
C2 M1 2 1 0 d d 0 d d
(14.67)
Note that most of the cross terms cancel. The drag coefficient breaks into several terms.

4
CD
=p
M1 2 1 !
2 Z 1 2 2 Z 1 2 Z 1 2 Z 1
A d B d B d
d + d d + d
C 0 d C 0 d C C 0 d C 0
(14.68)
The third term in (14.68) is

Z 1 Z 1
B d B B
d = d = ( (1) (0)) = 0. (14.69)
C C 0 d C C 0 C C

Finally

2 Z !
1 2 2 Z 1 2 2 Z 1
4 A d B d
CD
=p d + d + d .
M1 2 1 C 0 d C 0 d C 0
(14.70)
CHAPTER 14. THIN AIRFOIL THEORY 14-16

In the thin airfoil approximation, the drag is a sum of the drag due to thickness, drag due
to camber, and drag due to lift.

14.2 Similarity rules for high speed flight

The Figure below shows the flow past a thin symmetric airfoil at zero angle of attack. The
fluid is assumed to be inviscid and the flow Mach number, U11 /a1 where a1 2 = P1 /1
is the speed of sound, is assumed to be much less than one. The airfoil chord is c and the
maximum thickness is t1 . The subscript one is applied in anticipation of the fact that we
will shortly scale the airfoil to a new shape with subscript 2 and the same chord.

Figure 14.5: Pressure variation over a thin symmetric airfoil in low speed flow.

The surface pressure distribution is shown below the wing, expressed in terms of the pres-
sure coefficient.

Ps P1
C P1 = 1 2 (14.71)
2 1 U11

The pressure and flow speed throughout the flow satisfy the Bernoulli relation. Near the
airfoil surface

1 1
P1 + 1 U11 2 = Ps1 + 1 Us1 2 . (14.72)
2 2

The pressure is high at the leading edge where the flow stagnates, then as the flow ac-
celerates about the body, the pressure falls rapidly at first, then more slowly reaching a
minimum at the point of maximum thickness. From there the surface velocity decreases
CHAPTER 14. THIN AIRFOIL THEORY 14-17

and the pressure increases continuously to the trailing edge. In the absence of viscosity,
the flow is irrotational.

r u1 = 0 (14.73)

This permits the velocity to be described by a potential function.

u1 = r 1 (14.74)

When this is combined with the condition of incompressibility, r u1 = 0, the result is


Laplaces equation.

@2 1 @2 1
+ =0 (14.75)
@x1 2 @y1 2

Let the shape of the airfoil surface in (x1 , y1 ) be given by

y1 x
1
= 1 g . (14.76)
c c

where 1 = t1 /c is the thickness to chord ratio of the airfoil. The boundary conditions that
the velocity potential must satisfy are

@ 1 dy1 dg (x1 /c)
= U 11 = U 1 1
@y1 y1 =c1 g (
x1
) dx1 body d (x1 /c)
c
(14.77)

@ 1
! U 11 .
@x1 1

Any number of methods of solving for the velocity potential are available including the
use of complex variables. In the following we are going to restrict the airfoil to be thin,
1 << 1. In this context we will take the velocity potential to be a perturbation potential
so that

u 1 = U 11 + u 1 0
(14.78)
v1 = v1 0

or
CHAPTER 14. THIN AIRFOIL THEORY 14-18

@ 1
u1 = U 11 +
@x1
(14.79)
@ 1
v1 = .
@y1
The boundary conditions on the perturbation potential in the thin airfoil approximation
are

@ 1 dy1 dg (x1 /c)
= U 11 = U 1 1 1
@y1 y1 =0 dx1 body d (x1 /c)
(14.80)
@ 1
! 0.
@x1 1

The surface pressure coefficient in the thin airfoil approximation is



2 @ 1
C P1 = . (14.81)
U 11 @x1 y1 =0

Note that the boundary condition on the vertical velocity is now applied on the line y1 = 0.
In eect the airfoil has been replaced with a line of volume sources whose strength is
proportional to the local slope of the actual airfoil. This sort of approximation is really
unnecessary in the low Mach number limit but it is essential when the Mach number is
increased and compressibility eects come in to play. Equally, it is essential in this example
where we will map a compressible flow to the incompressible case.

14.2.1 Subsonic flow M1 < 1

Now imagine a second flow at a free-stream velocity, U12 in a new space (x2 , y2 ) over a new
airfoil of the same shape (defined by the function g (x/c) but with a new thickness ratio
2 = t2 /c << 1. Part of what we need to do is to determine how 1 and 1 are related to
one another. The boundary conditions that the new perturbation velocity potential must
satisfy are

@ 2 dy2 dg (x2 /c)
= U 12 = U 1 2 2
@y2 y2 =0 dx2 body d (x2 /c)
(14.82)
@ 2
! 0.
@x2 1
CHAPTER 14. THIN AIRFOIL THEORY 14-19

In this second flow the Mach number has been increased to the point where compressibility
eects begin to occur: the density begins to vary significantly and the pressure distribution
begins to deviate from the incompressible case. As long as the Mach number is not too
large and shock waves do not form, the flow will be nearly isentropic. In this instance the
2-D steady compressible flow equations are

@u2 @u2 1 @P
u2 + v2 + =0
@x2 @y2 @x2

@v2 @v2 1 @P
u2 + v2 + =0
@x2 @y2 @y2
(14.83)
@ @ @u2 @v2
u2 + v2 + + =0
@x2 @y2 @x2 @y2

P
= .
P1 1

Let

u 2 = U 12 + u 2 0

v2 = v2
(14.84)
= 1 + 0

P = P1 + P 0

where the primed quantities are assumed to be small compared to the free stream condi-
tions. When quadratic terms in the equations of motion are neglected the equations (14.83)
reduce to


1 U12 2 @u2 0 @v2 0
1 + = 0. (14.85)
P1 @x2 @y2

Introduce the perturbation velocity potential.

@ 2
u2 0 =
@x2
(14.86)
@ 2
v2 =
@y2
CHAPTER 14. THIN AIRFOIL THEORY 14-20

The equation governing the disturbance flow becomes

@2 2 @2 2
1 M1 2 2 + = 0. (14.87)
@x2 2 @y2 2

Notice that (14.87) is valid for both sub and supersonic flow in the thin airfoil approxi-
mation. Since the flow is isentropic, the pressure and velocity disturbances are related to
lowest order by

P 0 + 1 U12 u2 0 = 0 (14.88)

and the surface pressure coefficient retains the same basic form as in the incompressible
case.

2 @ 2
C P2 = (14.89)
U 12 @x2 y2 =0

Since we are at a finite Mach number, this last relation is valid only within the thin airfoil,
small disturbance approximation and therefore may be expected to be invalid near the
leading edge of the airfoil where the velocity change is of the order of the free stream
velocity. For example for the thin airfoil depicted in Figure 14.5 which is actually not
all that thin, the pressure coefficient is within 0.2 < Cp < 0.2 except over a very narrow
portion of the chord near the leading edge.
Equation (14.87) can be transformed to Laplaces equation, (14.75) using the following
change of variables

x2 = x1

1
y2 = q y1
1 M1 2 2 (14.90)

1 U 12
2 = 1
A U 11

where, at the moment, A is an arbitrary constant. The velocity potentials are related
by

1 U 12
2 (x2 , y2 ) = 1 (x1 , y1 ) (14.91)
A U 11
CHAPTER 14. THIN AIRFOIL THEORY 14-21

or
0 1

U 11 1
1 (x1 , y1 ) =A 2 @x 1 , q y1 A (14.92)
U 12 1 M1 2 2

and the boundary conditions transform as


0 1

@ 1 A2 A dg (x1 /c)
= U 11 @ q
@y1 y1 =0 1 M1 2 2 d (x1 /c)


@ 1 (14.93)
!0
@x1 1

@ 2
! 0.
@x2 1

The transformation between flows one and two is completed by the correspondence

A
1 = q 2 (14.94)
2
1 M1 2

or

t1 A t2
=q . (14.95)
c 1 M1 2 2 c

Finally the transformed pressure coefficient is

CP1 = ACP2 . (14.96)

These results may be stated as follows. The solution for incompressible flow over a thin
airfoil with shape g (x1 /c) and thickness ratio, t1 /c at velocity U1 is identical to the subsonic
compressible flow at velocity, U2 and Mach number M2 over an airfoil with a similar shape
but with the thickness ratio

q
t2 1 M 1 2 2 t1
= . (14.97)
c A c
CHAPTER 14. THIN AIRFOIL THEORY 14-22

The pressure coefficient for the compressible case is derived by adjusting the incompressible
coefficient using CP2 = CP1 /A. This result comprises several dierent similarity rules that
can be found in the aeronautical literature depending on the choice of the free constant A.
Perhaps the one of greatest interest is the Prandtl-Glauert rule that describes the variation
of pressure coefficient with Mach number for a body of a given shape and thickness ratio.
In this case we select

q
A= 1 M1 2 2 (14.98)

so that the two bodies being compared in (14.97) have the same shape and thickness ratio.
The pressure coefficient for the compressible flow is

C P1
C P2 = q . (14.99)
2
1 M1 2

Several scaled profiles are shown in Figure 14.6. Keep in mind the lack of validity of (14.99)
near the leading edge where the pressure coefficient is scaled to inaccurate values.

Figure 14.6: Pressure coefficient over the airfoil in Figure 14.5 at several Mach numbers
as estimated using the Prandtl-Glauert rule (14.99).

14.2.2 Supersonic similarity M1 > 1

All the theory developed in the previous section can be extended to the supersonic case
by simply replacing 1 M1 2 with M1 2 1. In this instance the mapping is between the
equation

@2 2 @2 2
M1 2 2 1 =0 (14.100)
@x2 2 @y2 2
CHAPTER 14. THIN AIRFOIL THEORY 14-23

and the simple wave equation

@2 1 @2 1
= 0. (14.101)
@x1 2 @y1 2

A generalized form of the pressure coefficient valid for subsonic and supersonic flow is
0 1
CP
=F@ q A (14.102)
A A 1 M1 2

where A is taken to be a function of 1 M1 2 .

14.2.3 Transonic similarity M1


=1

When the Mach number is close to one, the simple linearization used to obtain (14.87)
from (14.83) loses accuracy. In this case the equations (14.83) reduce to the nonlinear
equation

@u1 0 @v1 0 ( 1 + 1) M11 2 0 @u1 0


1 M1 1 2 + u1 = 0. (14.103)
@x1 @y1 U 11 @x1

In terms of the perturbation potential

@2 1 @2 1 ( 1 + 1) M11 2 @ 1 @ 2 1
1 M1 1 2 + = 0. (14.104)
@x1 2 @y1 2 U 11 @x1 @x1 2

This equation is invariant under the change of variables

x2 = x1
q
1 M1 1 2
y2 = q y1
(14.105)
1 M1 2 2

1 U 12
2 = 1
A U 11

where
CHAPTER 14. THIN AIRFOIL THEORY 14-24


1+ 2 1 M1 1 2 M1 2 2
A= . (14.106)
1+ 1 1 M1 2 2 M1 1 2

Notice that, due to the nonlinearity of the transonic equation (14.104) the constant A is
no longer arbitrary. The pressure coefficient becomes


1+ 2 1 M1 1 2 M1 2 2
C P1 = C P2 (14.107)
1+ 1 1 M1 2 2 M1 1 2

and the thickness ratios are related by

32
t2 1+ 1 1 M1 2 2 M1 1 2 t1
= . (14.108)
c 1+ 2 1 M1 1 2 M1 2 2 c

In the transonic case, it is not possible to compare the same body at dierent Mach numbers
or bodies with dierent thickness ratios at the same Mach number except by selecting gases
with dierent . For a given gas it is only possible to map the pressure distribution for
one airfoil to an airfoil with a dierent thickness ratio at a dierent Mach number. A
generalized form of (14.102) valid from subsonic to sonic to supersonic Mach numbers
is

1
0 1
CP ( + 1) M1 2 3
1 M1 2
2 =F@ 2
A. (14.109)
2
3
( + 1) M1 3

Prior to the advent of supercomputers capable of solving the equations of high speed flow,
similarity methods and wind tunnel correlations were the only tools available to the aircraft
designer and these methods played a key role in the early development of transonic and
supersonic flight.

14.3 Problems

Problem 1 - A thin, 2-D, airfoil is situated in a supersonic stream at Mach number M1


and a small angle of attack as shown in Figure 14.7.
The y-coordinate of the upper surface of the airfoil is given by the function

x x x
f (x) = A 1 (14.110)
C C C
CHAPTER 14. THIN AIRFOIL THEORY 25

Figure 14.7: Bi-convex airfoil at angle of attack

and the y-coordinate of the lower surface is

x x x
g (x) = A 1 (14.111)
C C C

where 2A/C << 1 and /C << 1. Determine the lift and drag coefficients of the air-
foil.

You might also like